Download Strategy for Nonenveloped Virus Entry

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Multi-state modeling of biomolecules wikipedia , lookup

SNARE (protein) wikipedia , lookup

Protein phosphorylation wikipedia , lookup

Thylakoid wikipedia , lookup

Magnesium transporter wikipedia , lookup

Cell membrane wikipedia , lookup

Protein wikipedia , lookup

G protein–coupled receptor wikipedia , lookup

Protein moonlighting wikipedia , lookup

Signal transduction wikipedia , lookup

Endomembrane system wikipedia , lookup

Nuclear magnetic resonance spectroscopy of proteins wikipedia , lookup

Intrinsically disordered proteins wikipedia , lookup

JADE1 wikipedia , lookup

List of types of proteins wikipedia , lookup

Protein purification wikipedia , lookup

Proteolysis wikipedia , lookup

Western blot wikipedia , lookup

Transcript
Strategy for Nonenveloped Virus Entry: a
Hydrophobic Conformer of the Reovirus
Membrane Penetration Protein µ1 Mediates
Membrane Disruption
Kartik Chandran, Diane L. Farsetta and Max L. Nibert
J. Virol. 2002, 76(19):9920. DOI:
10.1128/JVI.76.19.9920-9933.2002.
These include:
REFERENCES
CONTENT ALERTS
This article cites 70 articles, 41 of which can be accessed free
at: http://jvi.asm.org/content/76/19/9920#ref-list-1
Receive: RSS Feeds, eTOCs, free email alerts (when new
articles cite this article), more»
Information about commercial reprint orders: http://journals.asm.org/site/misc/reprints.xhtml
To subscribe to to another ASM Journal go to: http://journals.asm.org/site/subscriptions/
Downloaded from http://jvi.asm.org/ on February 27, 2014 by PENN STATE UNIV
Updated information and services can be found at:
http://jvi.asm.org/content/76/19/9920
JOURNAL OF VIROLOGY, Oct. 2002, p. 9920–9933
0022-538X/02/$04.00⫹0 DOI: 10.1128/JVI.76.19.9920–9933.2002
Copyright © 2002, American Society for Microbiology. All Rights Reserved.
Vol. 76, No. 19
Strategy for Nonenveloped Virus Entry: a Hydrophobic Conformer of
the Reovirus Membrane Penetration Protein ␮1 Mediates
Membrane Disruption†
Kartik Chandran,1 Diane L. Farsetta,2‡ and Max L. Nibert1*
Department of Microbiology and Molecular Genetics, Harvard Medical School, Boston, Massachusetts 02115,1 and
Cell and Molecular Biology Program, University of Wisconsin—Madison, Madison, Wisconsin 537062
The mechanisms employed by nonenveloped animal viruses to penetrate the membranes of their host cells
remain enigmatic. Membrane penetration by the nonenveloped mammalian reoviruses is believed to deliver a
partially uncoated, but still large (⬃70-nm), particle with active transcriptases for viral mRNA synthesis
directly into the cytoplasm. This process is likely initiated by a particle form that resembles infectious
subvirion particles (ISVPs), disassembly intermediates produced from virions by proteolytic uncoating. Consistent with that idea, ISVPs, but not virions, can induce disruption of membranes in vitro. Both activities
ascribed to ISVP-like particles, membrane disruption in vitro and membrane penetration within cells, are
linked to N-myristoylated outer-capsid protein ␮1, present in 600 copies at the surfaces of ISVPs. To understand how ␮1 fulfills its role as the reovirus penetration protein, we monitored changes in ISVPs during the
permeabilization of red blood cells induced by these particles. Hemolysis was preceded by a major structural
transition in ISVPs, characterized by conformational change in ␮1 and elution of fibrous attachment protein
␴1. The altered conformer of ␮1 was required for hemolysis and was markedly hydrophobic. The structural
transition in ISVPs was further accompanied by derepression of genome-dependent mRNA synthesis by the
particle-associated transcriptases. We propose a model for reovirus entry in which (i) primed and triggered
conformational changes, analogous to those in enveloped-virus fusion proteins, generate a hydrophobic ␮1
conformer capable of inserting into and disrupting cell membranes and (ii) activation of the viral particles for
membrane interaction and mRNA synthesis are concurrent events. Reoviruses provide an opportune system for
defining the molecular details of membrane penetration by a large nonenveloped animal virus.
The mechanisms employed by these viruses to cross the cellular membrane barrier are less well understood but may include
the formation of small protein-lined pores through the bilayer
analogous to those made by protein toxins such as colicin A
(55) or larger disruptions in bilayer integrity analogous to
those caused by membranolytic peptides such as magainin
(59). Nonenveloped viruses generally contain a capsid protein
or proteins that mediate membrane penetration. It has been
proposed that these “penetration proteins” also undergo
primed and triggered conformational transitions that allow
them to interact with the cellular membrane during entry. For
example, poliovirus virions incubated with receptor externalize
the N-myristoylated VP4 peptide and expose amphipathic sequences of VP1 that can insert into liposomes (5, 26). Adenovirus (Ad) virions incubated at low pH are thought to undergo
a conformational change in the penton base protein that renders the particles hydrophobic and capable of permeabilizing
liposomes (7, 58). As a step toward understanding the membrane penetration mechanism of the nonenveloped mammalian orthoreoviruses (reoviruses), we undertook studies to
identify and characterize entry-related changes in their capsid
proteins.
Reovirus virions are 85-nm particles comprising the doublestranded RNA genome enclosed by two concentric icosahedral
protein capsids (reviewed in reference 52) (Fig. 1A). The outer
capsid mediates delivery of viral particles into the cytoplasm of
host cells, where viral replication occurs. It has been proposed
that protein ␮1 (76 kDa, 200 trimers per particle), which ap-
Before viruses can launch their replicative programs, they
must first gain access to the interior of host cells where raw
materials and machinery essential for viral multiplication are
present. Enveloped animal viruses employ membrane fusion to
cross the membrane barrier and enter the cytoplasm, and the
viral surface proteins that mediate this process are termed
fusion proteins. These well-studied proteins, exemplified by
the influenza virus hemagglutinin (HA) and the tick-borne
encephalitis virus (TBEV) E protein (reviewed in references
33 and 63), share several features. (i) They are oligomeric
integral membrane proteins. (ii) They can adopt structurally
distinct water-soluble and membrane-seeking states. (iii) They
are primed, often by proteolytic cleavage, to undergo a conformational transition between these states. (iv) They are induced to undergo this transition by a triggering stimulus such
as low pH or receptor binding. The primed and triggered
conformational changes in the fusion proteins result in close
apposition of the viral and cellular bilayers, membrane merger,
and cytoplasmic delivery of the viral nucleocapsid(s).
Nonenveloped animal viruses, in contrast to their enveloped
counterparts, cannot utilize membrane fusion to enter cells.
* Corresponding author. Mailing address: Department of Microbiology and Molecular Genetics, Harvard Medical School, Boston, MA
02115. Phone: (617) 432-4829. Fax: (617) 738-7664. E-mail: mnibert
@hms.harvard.edu.
† This paper is dedicated to the memory of Lakshmi Chandran.
‡ Present address: ETAN Field Office, Social Justice Center, Madison, WI 53703.
9920
Downloaded from http://jvi.asm.org/ on February 27, 2014 by PENN STATE UNIV
Received 2 April 2002/Accepted 26 June 2002
VOL. 76, 2002
ENTRY-RELATED STRUCTURAL TRANSITION IN REOVIRUS ISVPs
pears to be found in virions mostly as autolytic fragments ␮1N
(4 kDa) and ␮1C (72 kDa) (53), participates directly in membrane penetration during entry into cells (16, 35, 42, 43, 51, 53).
␴3 (41 kDa, 600 monomers per particle), the major surface
protein of virions, interacts closely with the trimers of ␮1 (42,
61), thereby protecting them from the extracellular environment and regulating their activities. Protein ␴1 (50 kDa, 12
trimers per particle) forms fibers that extend from the fivefold
axes of virions and that mediate viral attachment to cellular
receptors (4, 27, 41). The fourth outer-capsid protein, ␭2 (144
kDa, 12 pentamers per particle), is involved in viral mRNA
capping and outer-capsid assembly but is not known to participate directly in entry (22, 45).
When virions are incubated with proteases in vitro, the outer
capsid is sequentially disassembled to yield two subvirion particles: infectious subvirion particles (ISVPs) and cores (reviewed in reference 52) (Fig. 1A). ISVPs lack ␴3 and contain
protein ␮1C as particle-bound fragments ␦ (59 kDa) and ␾ (13
kDa) (37, 51). Cores lack not only ␴3 but also ␮1 and its
fragments as well as ␴1 (37). Numerous observations suggest
that subvirion particles resembling the ISVP and core play
essential roles in reovirus infection. Proteolytic removal of ␴3
from an infecting virion is thought to prime the particle for
membrane penetration by freeing ␮1 to interact with a cellular
membrane (15, 42, 66) (also see below). Membrane penetration initiated by the ISVP-like particle is then believed to result
in cytoplasmic delivery of the primary transcriptase particle.
Like the core, this primary transcriptase particle is activated to
synthesize the viral mRNAs for translation and packaging (38,
62).
One line of evidence that supports the involvement of an
ISVP-like particle in membrane penetration is the capacity of
ISVPs, but not virions or cores, to promote permeabilization of
membranes in vitro. The target membranes include murine
L929 cell membranes (10), human and bovine red blood cell
(RBC) membranes (15, 16, 36), planar phospholipid bilayers
(67), and liposomes (K. Chandran and M. L. Nibert, unpublished data). In some of these previous experiments, Cs⫹ ions
were needed to promote the membrane interaction, for reasons that have remained largely mysterious but that may relate
to the capacity of Cs⫹ to promote reovirus uncoating (9, 11,
57). The M2 gene (which encodes ␮1) was identified as the
genetic determinant of a viral strain difference in the ISVPassociated permeabilization of L929 cell membranes (43), suggesting that ␮1 is a participant in this process. Also consistent
with a role for ␮1 in membrane permeabilization are observations that (i) ␮1 is modified with a myristoyl (C14 fatty-acyl)
group at its N terminus (53), (ii) viral particles containing
mutant forms of ␮1 show a reduced capacity to permeabilize
membranes (35), and (iii) in vitro addition of ␮1 to cores is
sufficient to restore the membrane-permeabilizing activity of
the particles (16). The crystal structure of the ␮1 trimer in
complex with three ␴3 monomers was recently determined at
2.8-Å resolution and suggested additional aspects of the functions of ␮1 in membrane penetration (42). Of particular note
for the present study was an indication that large conformational changes in the ␮1 trimer must be required for externalization of the N-myristoylated ␮1N peptide, which, it was proposed, inserts into the cellular membrane as a part of the
penetration mechanism (42, 53).
For a description of the molecular mechanism of reovirus
membrane penetration that can take advantage of the recently
determined ␮1 crystal structure (42), more information about
the biochemical and structural changes in ␮1 and other viral
proteins that accompany the membrane interactions is needed.
In the present study, we determined that ISVPs undergo a
major structural transition involving outer-capsid proteins ␮1,
Downloaded from http://jvi.asm.org/ on February 27, 2014 by PENN STATE UNIV
FIG. 1. Reovirus particles and their capacities to mediate hemolysis. (A) Surface views of the reovirus virion, ISVP, and core obtained
from transmission cryoelectron microscopy and three-dimensional image reconstruction as previously reported (22). Color coding of the
capsid proteins was applied for clarity (50). Capsid proteins are labeled
in a representative fashion. Bar, 20 nm. (B) Purified T3D virions,
ISVPs, or cores (1013 particles/ml) or an equal volume of virion buffer
lacking particles was incubated with bovine calf RBCs in virion buffer
for 15 min on ice or at 32°C. The amount of hemolysis induced by each
particle type is shown as the average ⫾ standard deviation from three
trials.
9921
9922
CHANDRAN ET AL.
␴1, and ␭2 that precedes, and is required for, virus-induced
membrane permeabilization and activation of the core-associated viral transcriptases. The nature of the structural changes
in ␮1 and ␴1 and their roles in membrane permeabilization by
viral particles were also investigated. Our results provide
mechanistic insights into two key steps in reovirus entry: membrane penetration and generation of the primary transcriptase
particles.
MATERIALS AND METHODS
solutions was measured from its molar extinction coefficient (⬃24,000 M⫺1
cm⫺1) just prior to use. Viral particles were incubated with bis-ANS (25 ␮M) at
the same reaction conditions used for hemolysis experiments (see above), except
that RBCs were not added. Reactions were initiated by transfer of samples to the
experimental temperature (ice or 32°C) and were terminated by their removal
onto ice. After resting on ice for at least 5 min, 10 to 20 ␮l of each sample was
diluted into virion buffer (100 ␮l total) in a 96-well black microplate designed for
fluorescence applications (Costar). bis-ANS fluorescence was measured at an
excitation wavelength (␭ex) of 405 nm and an emission wavelength (␭em) of 485
nm with the Spectramax Gemini XS fluorescence microplate reader (Molecular
Devices). An emission cutoff filter (␭cut ⬍ 475 nm) was employed to minimize
detection of scattered light from the excitation beam. bis-ANS fluorescence
intensity (I) was calculated from Icorrected ⫽ Isample ⫺ Iblank, where the blank
lacked viral particles but contained all other components of the reaction mixture.
Triton X-114 partitioning assay. Werck-Reichhart and coworkers’ (71) modification of Bordier’s original Triton X-114 partitioning protocol (8) was employed. Viral particles were diluted to a volume of 1 ml with virion buffer, chilled
for at least 5 min on ice, and incubated with ice-cold Triton X-114 (2% [vol/vol];
Sigma) for 1 h on ice with occasional mixing. Glycerol (40% [vol/vol]) was then
added. Samples were incubated for 10 min at 37°C, followed by centrifugation
(500 ⫻ g, 5 min, room temperature) to induce separation into detergent-poor
and -rich phases. Inclusion of glycerol in the partitioning mixture allowed flotation, rather than sedimentation, of the detergent-rich phase (71), thereby removing the concern that the detergent-rich fraction contained particles merely due to
pelleting during centrifugation. The bottom of each centrifuge tube was punctured with a needle, and both detergent-poor and -rich phases were harvested by
collecting drops. The detergent-poor fraction was reextracted with fresh Triton
X-114 and glycerol as described above. Detergent-rich fractions from both extractions were pooled, and the second detergent-poor fraction was discarded.
Viral proteins within the detergent-poor and -rich fractions were precipitated
with trichloroacetic acid (TCA) as follows. Samples were incubated with TCA
(10% [wt/vol]) and purified carbonic anhydrase (5 ␮g) as a carrier protein for at
least 1 h on ice. Precipitated material was concentrated by centrifugation (10,000
⫻ g for 30 min at 4°C), washed twice with ice-cold acetone, and dried under
reduced pressure. Dried pellets were solubilized in room-temperature virion
buffer. Following addition of Laemmli sample buffer, samples were disrupted by
boiling for 5 min. Viral proteins were resolved by SDS-PAGE.
RESULTS
T3D ISVPs, but not virions or cores, induce lysis of RBCs.
Previous work showed that only ISVPs can release 51Cr from
preloaded L929 cells (10, 43), form ion-permeant channels in
planar phospholipid bilayers (67), and lyse RBCs in the presence of Cs⫹ ions (15, 16, 36). To test whether certain reovirus
particles can permeabilize RBC membranes in a Cs⫹-free
buffer, we incubated purified T3D virions, ISVPs, or cores (Fig.
1A) with human RBCs either on ice or at 32°C and then
measured the extent of hemolysis. None of the particles induced hemolysis upon incubation on ice (Fig. 1B), but, when
incubated with RBCs at 32°C, T3D ISVPs lysed essentially all
of the cells (Fig. 1B). In contrast, T3D virions and cores were
inactive for hemolysis at these and all conditions tested (Fig.
1B; data not shown). Thus ISVPs, but not virions or cores, can
induce permeabilization of different types of membranes, both
in vitro and in cell culture, in the presence or absence of Cs⫹
ions. In addition, the failure of ISVPs to induce hemolysis at
4°C suggests that the permeabilization process is temperature
dependent and may require changes in protein conformation
and/or lipid mobility.
Genome segment M2, which encodes ␮1, determines a difference in hemolytic capacity between T1L and T3D ISVPs. To
extend our findings to particles of other reovirus strains, we
tested the capacity of T1L ISVPs to mediate hemolysis in a
Cs⫹-free buffer. T1L ISVPs failed to induce RBC lysis upon
incubation at the conditions used with T3D ISVPs above (data
Downloaded from http://jvi.asm.org/ on February 27, 2014 by PENN STATE UNIV
Cells. Spinner-adapted murine L929 cells were grown in Joklik’s modified
minimal essential medium (Irvine Scientific Co., Irvine, Calif.) supplemented to
contain 2% fetal bovine serum and 2% bovine calf serum (HyClone Laboratories, Logan, Utah) in addition to 2 mM glutamine, 100 U of penicillin/ml, and 100
␮g of streptomycin/ml (Irvine Scientific Co.).
Virions, ISVPs, and cores. Virions of reoviruses T1L, T3D, and the T1L ⫻
T3D reassortants in Fig. 2 were obtained by the standard protocol (27) and
stored in virion buffer (150 mM NaCl, 10 mM MgCl2, 10 mM Tris [pH 7.5]).
Purified T1L ISVPs were obtained by the same protocol. Unless mentioned
otherwise, nonpurified ISVPs were obtained by digesting virions at a concentration of 1013 particles/ml with N␣-p-tosyl-L-lysine chloromethyl ketone-treated
chymotrypsin (200 ␮g/ml) for 8 to 20 min at 32°C. Digestion was stopped by
addition of ethanolic phenylmethylsulfonyl fluoride (2 to 5 mM). Purified cores
were obtained from virions as described previously (54). Particle concentrations
were estimated on the basis of A260 (20).
SDS-PAGE and immunoblotting. Sodium dodecyl sulfate-polyacrylamide gel
electrophoresis (SDS-PAGE) was carried out on 10% acrylamide gels as described previously (51). Viral proteins were detected by staining with Coomassie
brilliant blue R-250 (Sigma) or were transferred to a nitrocellulose membrane
with the Mini-TransBlot kit (Bio-Rad, Hercules, Calif.) and visualized by immunoblotting. The ␮1-specific mouse monoclonal antibody 10H2 (70) and T1L
␴1-specific rabbit antisera raised against a glutathione S-transferase–␴1 fusion
protein (to be described elsewhere) were employed at a 1:1,000 dilution as
primary antibodies. Mouse- or rabbit-specific goat immunoglobulin Gs conjugated to alkaline phosphatase (Sigma) were used at a 1:2,000 dilution as secondary antibodies. Antibody binding was detected with colorimetric reagents
p-nitroblue tetrazolium chloride and 5-bromo-4-chloro-3-indolylphosphate ptoluidine salt (Bio-Rad) according to the manufacturer’s instructions.
Hemolysis. Citrated bovine calf RBCs (Colorado Serum Co., Denver, Colo.)
were washed with phosphate-buffered saline supplemented with 2 mM MgCl2
(PBS-Mg2⫹) and suspended in PBS-Mg2⫹ at a stock concentration of 30%
(vol/vol) just prior to use. All experiments were performed in a 4°C cold room to
minimize variability in results. Viral particles were incubated with RBCs (3%
[vol/vol]) in virion buffer or hemolysis reaction buffer (50 mM Tris-Cl [pH 7.5])
in a total volume of 20 to 100 ␮l. Na⫹ or Cs⫹ ions were added to hemolysis
reaction mixtures as their chloride salts from stock solutions prepared in water.
Reactions were initiated by transfer of samples to the experimental temperature
(ice or 32°C) and terminated by their removal onto ice. After resting on ice for
at least 5 min, cells within each sample were pelleted by centrifugation (300 ⫻ g
for 5 min at 4°C), and 10 to 20 ␮l of the supernatant was diluted into virion buffer
(100 ␮l total) in a 96-well microplate (Costar, Cambridge, Mass.). The amount
of hemoglobin released from RBCs was quantitated by measuring A415 with a
microplate reader (Molecular Devices, Sunnyvale, Calif.) and by using the equation percent hemolysis ⫽ {[A415 (sample) ⫺ A415 (blank)]/[A415 (detergent) ⫺ A415
(blank)]} ⫻ 100%, where the blank lacked viral particles but contained all other
components of the reaction mixture and the detergent was either Triton X-100
(Sigma; 1% [vol/vol]) or Nonidet P-40 (Sigma; 0.5% [vol/vol]). Addition of either
detergent to the hemolysis reaction mixture effected complete RBC lysis.
Proteolysis assay for conformational changes in viral proteins. Viral particles
were chilled on ice for at least 5 min and incubated with trypsin (100 ␮g/ml) for
20 to 60 min on ice. Reactions were stopped by addition of soybean trypsin
inhibitor (300 ␮g/ml) and further incubation for 5 to 10 min on ice. Incubation
of samples containing viral particles with trypsin for longer times (up to 2 h
tested; data not shown) did not appear to alter the pattern of digestion observed.
Following addition of Laemmli sample buffer, samples were disrupted by boiling
for 2 to 5 min and were subjected to SDS-PAGE.
bis-ANS fluorescence. High-purity bis-ANS (4,4⬘-dianilino-1,1⬘-binaphthyl5,5⬘-disulfonic acid, dipotassium salt) was obtained from Molecular Probes
(Junction City, Oreg.). bis-ANS stock solutions (5 to 10 mM) were prepared in
methanol and stored in the dark at 4°C. The concentration of bis-ANS in these
J. VIROL.
VOL. 76, 2002
ENTRY-RELATED STRUCTURAL TRANSITION IN REOVIRUS ISVPs
9923
not shown; Fig. 2A). To determine whether this difference
between T1L and T3D ISVPs arises from one or more genetic
differences between the two strains, we analyzed the behavior
of a panel of T1L ⫻ T3D reassortant strains in the hemolysis
assay. The capacity of ISVPs to induce hemolysis segregated
with the parental origin of their M2 genome segments (Fig.
2C). All reassortant ISVPs containing an M2 segment derived
from T3D were positive for hemolysis, whereas all containing
an M2 segment derived from T1L were negative. No other
genome segment contributed significantly to the phenotypic
difference between T1L and T3D ISVPs (Fig. 2C). LuciaJandris and coworkers (43) previously found that M2 also
Downloaded from http://jvi.asm.org/ on February 27, 2014 by PENN STATE UNIV
FIG. 2. Conformational state of the ␮1 protein within hemolysis reactions and reassortant genetic analysis of the difference in hemolysis
between T1L and T3D ISVPs. (A) Nonpurified ISVPs of strains T1L (circles), T3D (squares), E3 (diamonds), or EB144 (triangles) (8 ⫻ 1012
particles/ml) were incubated with RBCs in virion buffer for different times at 32°C. Only 0- and 20-min time points were taken for EB144 and E3
ISVPs. Each sample was divided into two equal aliquots. One aliquot was used to measure the amount of hemolysis. Results from two trials are
shown. (B) The second aliquot was treated with trypsin for 30 min on ice, and digestion was stopped by addition of soybean trypsin inhibitor.
Samples were subjected to SDS-PAGE, and ␦, the major fragment of ␮1 within ISVPs, was visualized by immunoblotting with ␮1-specific
monoclonal antibody 10H2. The positions of ␦ and an unidentified proteolytic fragment derived from ␦ (ⴱ) are marked. (C) Reassortant strains
chosen for this study represent all 16 possible combinations of T1L (L) and T3D (D) alleles for genome segments L2, M2, S1, and S4, which encode
outer-capsid proteins ␭2, ␮1, ␴1, and ␴3, respectively. Strains are listed in order of decreasing amount of hemolysis. The allelic origin of genome
segment M2 is in boldface. For determining the strain behaviors, nonpurified ISVPs (1013 particles/ml) were incubated with RBCs in virion buffer
for 7 min at 32°C. Each sample was divided into two equal aliquots. One aliquot was used to measure the amount of hemolysis. Averages from
five or six trials (T1L and T3D) and one or two trials (reassortant strains) are shown. The second aliquot was treated with trypsin for 20 min on
ice, and digestion was stopped by addition of soybean trypsin inhibitor. Samples were subjected to SDS-PAGE, and the ␦ fragment of ␮1 was
visualized by Coomassie staining. Conformational (conf.) change: ⫹ or ⫺, loss or retention of ␦ after trypsin treatment, respectively; nd, not
determined.
9924
CHANDRAN ET AL.
FIG. 3. Effect of Cs⫹ ions and RBCs on hemolysis and the conformational state of ␮1. (A and B) Nonpurified ISVPs of strains T1L
(circles) or T3D (squares) (4 ⫻ 1012 particles/ml) were incubated with
RBCs in reaction buffer (50 mM Tris-Cl [pH 7.5]) containing CsCl
(300 mM) for different times at 32°C. Measurement of the amount of
hemolysis (A) and assessment of the protease sensitivity of ␦ (B) were
carried out as described for Fig. 2A and B. Results from two trials are
shown in panel A. (C) Samples were generated as described for panels
A and B except that no RBCs were added. Samples were treated with
trypsin for 30 min on ice, and digestion was stopped by addition of
soybean trypsin inhibitor. Samples were subjected to SDS-PAGE, and
the viral proteins were visualized by Coomassie staining.
ions in the hemolysis reaction. Accordingly, we repeated the
hemolysis experiments in a low-Na⫹ buffer supplemented with
K⫹ or Cs⫹ ions. Consistent with our hypothesis, the ␮1 protein
in T1L ISVPs became protease sensitive upon incubation with
RBCs at 32°C in the presence of K⫹ (data not shown) or Cs⫹
ions (Fig. 3B), indicating that T1L ␮1 had indeed changed
conformation. Moreover, T1L ISVPs induced hemolysis at
these conditions (Fig. 3A), consistent with previous evidence
that they have that capacity (15, 16, 36). In addition, the presence of Cs⫹ in the hemolysis reaction decreased the onset time
of ␮1 conformational change with T3D ISVPs (compare Fig.
Downloaded from http://jvi.asm.org/ on February 27, 2014 by PENN STATE UNIV
genetically controls the difference between T1L and T3D ISVPs in their capacity to release 51Cr from preloaded L929 cells
(T1L ISVPs were inactive in that assay as well). The simplest
hypothesis based on these results is that ␮1, the protein encoded by M2 and the major surface protein of ISVPs (Fig. 1A),
participates directly in permeabilization of both RBC and L929
cell membranes. A related hypothesis is that the ␮1 proteins
encoded by the T1L and T3D alleles of M2 must differ in their
capacities to induce permeabilization at the assay conditions.
Protein ␮1 changes conformation in association with hemolysis. We used sensitivity to protease digestion to monitor
conformational changes in viral particles during hemolysis reactions. T1L or T3D ISVPs were incubated with RBCs at 32°C
for different times. At each time point, samples were removed
to ice and divided into two aliquots. One aliquot was analyzed
for RBC lysis. The other was incubated with trypsin for 60 min
on ice. Viral proteins in the trypsin-treated samples were resolved by SDS-PAGE and detected by immunoblotting with
protein-specific antibodies. Evidence for a change in protease
sensitivity over time was obtained only in the case of T3D ␮1
(Fig. 2B; data not shown for the other viral proteins, but see
Fig. 3C). Specifically, the ␦ fragment of ␮1 in T3D ISVPs
became sensitive to trypsin cleavage after between 8 and 12
min of incubation at 32°C. This corresponded to the onset time
of hemolysis induced by T3D ISVPs (Fig. 2A). When we repeated the experiment with T1L ISVPs, which did not mediate
hemolysis under these conditions (Fig. 2A), we found that T1L
␮1 remained resistant to cleavage by trypsin over the entire
time course (Fig. 2B). Essentially identical results were obtained with several other proteases (data not shown). These
findings indicate that protein ␮1 acquires a protease-sensitive
conformation when T3D ISVPs are incubated with RBCs at
32°C, but not at 4°C, and that this conformational change in ␮1
occurs in temporal correlation with hemolysis.
We also tested ISVPs derived from T1L ⫻ T3D reassortant
strains in the protease sensitivity assay. Reassortant ISVPs that
scored positive for hemolysis (i.e., those that carry T3D M2)
underwent ␮1 conformational change in association with hemolysis, whereas reassortant ISVPs that scored negative for
hemolysis (i.e., those that carry T1L M2) did not undergo ␮1
conformational change detected by protease treatment (Fig.
2B and C). These results provide additional evidence that ␮1
changes conformation in association with hemolysis. They also
show that the difference between T1L and T3D ISVPs in the
propensity of ␮1 to change conformation is determined by one
or more sequence differences between the T1L and T3D alleles
of M2.
Rates of hemolysis and ␮1 conformational change increase
in parallel in the presence of Csⴙ ions. T3D ISVPs are proteolytically digested to cores more efficiently when incubated
with K⫹, Rb⫹, or Cs⫹ ions in place of Na⫹ or Li⫹ ions (9, 11).
Since differences among viral strains in the susceptibility of
ISVPs to undergo proteolytic digestion to cores are genetically
determined by M2, encoding ␮1 (10), these results suggest that
K⫹, Rb⫹, and Cs⫹ ions allow more efficient conversion of
ISVPs to cores by increasing the capacity of ␮1 to undergo
protease digestion (12, 13). Based on these previous findings,
we speculated that the propensity of T1L ␮1 to acquire a
protease-sensitive conformation when ISVPs are incubated
with RBCs may be enhanced by inclusion of K⫹, Rb⫹, or Cs⫹
J. VIROL.
VOL. 76, 2002
ENTRY-RELATED STRUCTURAL TRANSITION IN REOVIRUS ISVPs
FIG. 4. Requirement of ␮1 conformational change for hemolysis.
(A) Schematic diagram for the experiment. Nonpurified T1L ISVPs (4
⫻ 1012 particles/ml) were incubated in reaction buffer containing CsCl
(300 mM) for different times at 32°C. After a 5-min incubation on ice,
RBCs were added to each sample, and hemolysis reactions were allowed to proceed for 15 min on ice. (B) The amount of hemolysis in
each sample was determined. Results from two trials are shown.
(C) Same as panel B except that trypsin was added to samples instead
of RBCs. Protease digestion was allowed to proceed for 60 min on ice
and was stopped by addition of soybean trypsin inhibitor. Samples
were subjected to SDS-PAGE, and the viral proteins were visualized
by Coomassie staining.
RBCs were then added to all four aliquots, and hemolysis
reactions were allowed to proceed on ice (Fig. 5B). Parallel
reaction mixtures were subjected to SDS-PAGE and Coomassie staining to visualize the viral proteins (Fig. 5C). As expected, aliquot 1 was wholly competent for hemolysis. However, aliquot 2 was wholly deficient for hemolysis. Comparison
of the protein compositions of viral particles within aliquots 1
and 2 revealed that loss of hemolytic capacity correlated with
a substantial reduction in the complement of ␮1 in aliquot 2.
The mere presence of extra proteins (i.e., trypsin and soybean
trypsin inhibitor) did not explain the inhibition of hemolysis,
since inactivated trypsin in aliquot 3 had little or no effect on
either the extent of hemolysis or ␮1 content (aliquot 3). Also
Downloaded from http://jvi.asm.org/ on February 27, 2014 by PENN STATE UNIV
3A and 2A), and for both T1L and T3D ISVPs, there was a
correspondence between the onset times of ␮1 conformational
change and hemolysis (Fig. 3A). Thus, Cs⫹ ions accelerate in
parallel both ␮1 conformational change and hemolysis. The
consistent temporal correlation between these phenomena argues that they are mechanistically linked.
RBCs are dispensable for ␮1 conformational change. In the
experiments described thus far (Fig. 2 and 3), ␮1 conformational change was observed in hemolysis reactions with RBCs.
To test whether interaction of viral particles with one or more
components of the RBC membrane are required for induction
of ␮1 conformational change, we repeated the experiments
described above but omitted RBCs from the samples. Protein
␮1 acquired a protease-sensitive conformation when either
T1L or T3D ISVPs were incubated with Cs⫹ ions at 32°C in the
absence of RBCs (Fig. 3C). Moreover, the kinetics of ␮1 conformational change in the absence or presence of RBCs were
essentially indistinguishable (Fig. 3C). RBCs also had little or
no effect on the rate of ␮1 conformational change in experiments carried out in the presence of Na⫹ ions (data not
shown). Therefore, induction and maintenance of a proteasesensitive conformation of ␮1 do not require interaction of viral
particles with a component of RBC membranes.
␮1 conformational change is required for hemolysis. We
investigated the causal relationship between ␮1 conformational change and hemolysis by attempting to decouple the two
events. T1L ISVPs were incubated for different times at 32°C
in the absence of RBCs and then removed to ice. RBCs were
then added, and hemolysis reactions were allowed to proceed
on ice (see Fig. 4A for a schematic diagram). The conformational status of ␮1 after the incubation at 32°C was assessed by
protease treatment of parallel samples on ice. The only samples in which hemolysis occurred at 4°C were those containing
a protease-sensitive conformer of ␮1 generated by preincubation of the ISVPs at 32°C (Fig. 4B and C). These results
strongly suggest that ␮1 conformational change is required for
hemolysis. They also suggest that hemolysis is normally
blocked at 4°C because the conformational change cannot occur at that temperature. The RBC lysis step(s) is, by comparison, less dependent on temperature: hemolysis mediated by
samples containing a protease-sensitive conformer of ␮1 was
slowed from a few seconds at 32°C to ⬃5 min at 4°C but
occurred efficiently at both temperatures (data not shown).
A protease-sensitive conformer of ␮1 is necessary for hemolysis. Although the preceding experiments suggested that ␮1
conformational change is required for hemolysis (Fig. 2 to 4),
it remained possible that the conformationally altered ␮1 protein per se is dispensable. For instance, hemolysis might be
carried out by another viral protein activated by ␮1 conformational change. To distinguish between these possibilities, we
used a modified form of the order-of-incubation experiment
described in the legend for Fig. 4 (see Fig. 5A for a schematic
diagram). T1L ISVPs were allowed to undergo the ␮1 conformational change in the absence of RBCs by incubation with
Cs⫹ ions at 32°C and then removed to ice. The chilled sample
was divided into four aliquots. Aliquots 1 and 4 were left
untreated. Aliquot 2 was incubated with trypsin on ice, after
which the protease was inactivated with soybean trypsin inhibitor. Aliquot 3 was treated with inactivated trypsin generated
by preincubation of protease with soybean trypsin inhibitor.
9925
9926
CHANDRAN ET AL.
J. VIROL.
ruled out was the possibility that trypsin inhibits hemolysis by
acting on RBCs rather than viral particles; RBCs pretreated
with trypsin were lysed efficiently (aliquot 4). These findings
strongly suggest that a protease-sensitive conformer of ␮1 is
necessary for hemolysis.
Hydrophobicity of viral particles increases in concert with
␮1 conformational change. Hemolysis induced by ISVPs almost certainly involves the interaction of viral protein regions
with the RBC membrane bilayer. Since change in ␮1 conformation is required for hemolysis, we suspected that this change
may expose hydrophobic sequences previously buried within
ISVPs. In the following experiments, we used two different
assays to detect changes in the hydrophobicity of viral particles:
fluorescence of the polarity-sensitive dye bis-ANS and partitioning of particles into detergent Triton X-114.
bis-ANS fluorescence. bis-ANS is essentially nonfluorescent
in aqueous solution but becomes strongly fluorescent when
dissolved in nonpolar solvents or when bound to surface-accessible hydrophobic sites within proteins (46). T1L ISVPs
were incubated with bis-ANS and either Na⫹ or Cs⫹ ions for
different times at 32°C and then removed to ice. Fluorescence
in each sample was measured in a microplate reader. Parallel
FIG. 6. Effect of ␮1 conformational change on bis-ANS fluorescence. (A) Nonpurified T1L ISVPs (4 ⫻ 1012 particles/ml) were incubated in reaction buffer containing bis-ANS (25 ␮m) and NaCl
(squares) or CsCl (circles) (300 mM) for different times at 32°C. The
amount of bis-ANS fluorescence was then measured. Results from two
trials are shown. The conformational state of ␮1 in each sample was
assessed by trypsin treatment and is indicated below the graph. ⫹ and
⫺, loss and retention of ␦ after trypsin treatment, respectively. ␮1ⴱ,
protease-sensitive conformer of ␮1. (B) Schematic diagram for the
order-of-incubation experiment. The experiment was performed as for
Fig. 5 except that bis-ANS (25 ␮M) was added to samples instead of
RBCs. (C) The amount of bis-ANS fluorescence in each sample from
panel B was determined. Results from two trials are shown as superimposed open and filled bars.
samples were exposed to trypsin on ice to assess the conformational status of ␮1 (data not shown). Little or no change in
bis-ANS fluorescence was seen when T1L ISVPs were incubated with Na⫹ ions (i.e., at conditions that did not induce ␮1
conformational change over the entire time course) (Fig. 6A).
In contrast, when T1L ISVPs were incubated with Cs⫹ ions, a
large increase in bis-ANS fluorescence was observed. The onset time of this increase corresponded to that of ␮1 conformational change (Fig. 6A). Experiments with T3D ISVPs in the
Downloaded from http://jvi.asm.org/ on February 27, 2014 by PENN STATE UNIV
FIG. 5. Requirement of a protease-sensitive ␮1 conformer for hemolysis. (A) Schematic diagram for the experiment. Nonpurified T1L
ISVPs (4 ⫻ 1012 particles/ml) were incubated in reaction buffer containing CsCl (300 mM) for 8 min at 32°C and then removed to ice. The
sample was divided into four aliquots, and each was treated as indicated. ISVPⴱ, particle type derived from ISVPs that have undergone
␮1 conformational change (see text for more information); try, trypsin;
sbti, soybean trypsin inhibitor; try†, trypsin (1 mg/ml) pretreated with
soybean trypsin inhibitor (3 mg/ml) for 20 min on ice; rbc†, RBCs
pretreated first with trypsin for 20 min on ice and then with soybean
trypsin inhibitor for 10 min. (B) The amount of hemolysis in each
sample was determined. Averages ⫾ standard deviations from three
trials are shown. (C) Same as panel B except that Laemmli sample
buffer was added instead of RBCs. Samples were boiled and subjected
to SDS-PAGE, and the viral proteins were visualized by Coomassie
staining.
VOL. 76, 2002
ENTRY-RELATED STRUCTURAL TRANSITION IN REOVIRUS ISVPs
presence of Na⫹ ions confirmed that bis-ANS fluorescence
increased in concert with the acquisition of a protease-sensitive
conformation by protein ␮1 (data not shown). These findings
suggest that viral particles expose hydrophobic sequences to
the aqueous milieu in association and temporal correlation
with ␮1 conformational change.
Triton X-114 partitioning. Nonionic detergent Triton X-114
forms a homogeneous phase in aqueous solution at 4°C but
separates into discrete detergent-rich and detergent-poor
phases as the temperature is elevated past 20°C (8). Proteins
that form mixed micelles with detergent by virtue of their
exposed hydrophobic regions (e.g., integral membrane proteins) partition into the detergent-rich phase, whereas proteins
lacking extensive hydrophobic surfaces (e.g., many cytosolic
proteins) partition into the detergent-poor phase. T1L ISVPs
were incubated with Na⫹ or Cs⫹ ions at 32°C and removed to
ice. Triton X-114 was then added to the samples. After further
incubation on ice, detergent-rich and -poor phases were separated by incubation at 37°C, and the detergent-rich phase was
floated by centrifugation. Both phases were harvested, and the
distribution of viral proteins in each fraction was examined by
SDS-PAGE and Coomassie staining. When T1L ISVPs were
incubated with Na⫹ ions at 32°C prior to Triton X-114 extraction, viral proteins were predominantly located in the detergent-poor fraction, indicating that most particles had partitioned into the detergent-poor phase (Fig. 7A). In contrast,
when T1L ISVPs were incubated with Cs⫹ ions at 32°C before
Triton X-114 extraction, viral proteins were predominantly
located in the detergent-rich fraction, indicating that most particles had shifted into the detergent-rich phase (Fig. 7A). Tryp-
sin treatment of aliquots removed before Triton X-114 extraction indicated that, as expected, ␮1 had acquired a proteasesensitive conformation in particles incubated at 32°C with Cs⫹
ions but not Na⫹ ions (data not shown). These findings, together with those with bis-ANS (Fig. 6A), strongly suggest that
␮1 conformational change causes hydrophobic protein regions
to be exposed at the surfaces of viral particles.
␮1 exposes hydrophobic regions as a result of its conformational change. The preceding data make ␮1 a prime candidate
to be the viral protein that exposes hydrophobic regions preceding hemolysis. To test that hypothesis, we performed an
order-of-incubation experiment exactly as described for Fig. 5,
with the exception that bis-ANS was added to the reaction
mixtures instead of RBCs (see Fig. 6B for a schematic diagram). T1L ISVPs were incubated with Cs⫹ ions at 32°C to
allow the ␮1 conformational change to occur and then removed to ice. Reaction mixtures were split into three aliquots.
Aliquot 1 was left untreated, aliquot 2 was treated with trypsin,
and aliquot 3 was treated with inactivated trypsin. bis-ANS was
then added to the reaction mixtures, and the fluorescence in
each was measured after a further incubation on ice (Fig. 6C).
bis-ANS fluorescence in aliquot 2 was substantially reduced
(by ⬃80%) relative to that in aliquots 1 and 3. Comparison of
the protein compositions of the three samples revealed that
loss of fluorescence correlated with a substantial reduction in
the complement of ␮1 in aliquot 2 (data not shown). The
findings suggest that ␮1 exposes previously buried hydrophobic
regions as a result of its conformational change.
Receptor-binding protein ␴1 is lost from particles in concert with ␮1 conformational change. Closer examination of
Coomassie-stained gels from the Triton X-114 experiment revealed that, although most viral proteins from the particles
preincubated with Cs⫹ ions were predominantly located in the
detergent-rich fraction, a 50,000-Mr band migrating at the expected position of receptor-binding protein ␴1 appeared to be
enriched in the detergent-poor fraction (Fig. 7A). Immunoblot
analysis of detergent-rich and -poor fractions with a ␴1-specific
antibody confirmed that most ␴1 proteins did not partition into
the detergent-rich phase along with other viral proteins (Fig.
7B). This observation suggests that ␮1 conformational change
is associated with the loss of ␴1 from viral particles and that the
hydrophobicity of ␴1 does not increase substantially in conjunction with its elution from particles.
To test further whether viral particles that had undergone
␮1 conformational change had also lost ␴1, we attempted to
purify those particles. T1L ISVPs were incubated with Na⫹ or
Cs⫹ ions at 32°C and then transferred to ice. Trypsin treatment
of aliquots removed prior to purification confirmed that ␮1
had acquired a protease-sensitive conformation in the viral
particles incubated at 32°C with Cs⫹ ions but not Na⫹ ions
(data not shown). Remaining samples were loaded atop preformed CsCl gradients, and viral particles (density [␳] ⬇ 1.38
g/cm3) were separated from particle-free proteins (␳ ⬇ 1.30
g/cm3) according to density by ultracentrifugation. Visual inspection of the gradients revealed that viral particles incubated
with Na⫹ ions before centrifugation had formed a homogeneous band with an opalescent appearance, as typically seen
when ISVPs are concentrated on CsCl density gradients (Fig.
8A). In contrast, particles incubated with Cs⫹ ions before
centrifugation produced an inhomogeneous band composed of
Downloaded from http://jvi.asm.org/ on February 27, 2014 by PENN STATE UNIV
FIG. 7. Effect of ␮1 conformational change on Triton X-114 partitioning of viral proteins. Nonpurified T1L ISVPs (4 ⫻ 1011 particles/
ml) were incubated in reaction buffer containing NaCl or CsCl (200
mM) for 40 min at 37°C and then removed to ice. Samples were
extracted with Triton X-114 as described in Materials and Methods.
The detergent-poor (aq) and detergent-rich (det) fractions were subjected to SDS-PAGE, followed by Coomassie staining to visualize the
viral proteins (A) or immunoblot analysis with a T1L ␴1-specific polyclonal antiserum (B). ⴱ, position of an unidentified proteolytic fragment that comigrated with ␴1. The conformational state of ␮1 in each
fraction was assessed by trypsin treatment and is indicated below the
gel.
9927
9928
CHANDRAN ET AL.
white flocculated material (Fig. 8A). A flocculent band was
also obtained with T3D ISVPs that had been incubated with
Na⫹ ions at 32°C (data not shown). The opalescent and flocculent bands migrated at similar positions in CsCl gradients,
indicating that the particles in them had similar densities (data
not shown). The change in morphology of the particle band
from opalescent to flocculent suggested that viral particles that
had undergone ␮1 conformational change were prone to ag-
gregation at high concentrations. This agrees with our finding
that ␮1 conformational change enhances the hydrophobicity of
particles (Fig. 6 and 7). To determine the protein compositions
of viral particles in the opalescent and flocculent bands, the
samples were subjected to SDS-PAGE and monitored by Coomassie staining or immunoblot analysis with a ␴1-specific antibody. As expected, particles in the opalescent band closely
resembled ISVPs; the resemblance extended to the presence of
␮1 fragment ␦ and protein ␴1 (Fig. 8B). Particles comprising
the flocculent band also resembled ISVPs in containing an
approximately full complement of the ␦ fragment of ␮1; however, they contained little or no ␴1 (Fig. 8B). These findings
confirmed that viral particles that had undergone ␮1 conformational change had also lost ␴1. They also indicated that the
central ␦ fragment of ␮1 remained associated with particles
even after ␮1 acquired a protease-sensitive conformation.
We next sought to define the temporal relationship between
␮1 conformational change and loss of ␴1 from particles. T1L
ISVPs were incubated with Na⫹ or Cs⫹ ions for different times
at 32°C and then removed to ice. Each sample was loaded onto
a sucrose cushion, and viral particles were pelleted by ultracentrifugation. A portion of supernatant from the top of each
tube (Fig. 8C) and the particle pellet at the bottom of each
tube (data not shown) were subjected to SDS-PAGE and immunoblot analysis with a ␴1-specific antibody. Particles incubated with Na⫹ ions did not undergo ␮1 conformational
change and released little or no ␴1 over the time course. In
contrast, particles incubated with Cs⫹ ions underwent ␮1 conformational change and released most or all of their ␴1 fibers,
with similar onset times for both structural changes. These
findings indicate that ␴1 is lost from viral particles in concert
with acquisition of the protease-sensitive conformation of ␮1.
Viral particles that contain a protease-sensitive ␮1 conformer and lack ␴1 are activated for mRNA synthesis. Both
virions and ISVPs are inactive with regard to genome-dependent synthesis of viral mRNAs. In contrast, cores, which lack
outer-capsid proteins ␮1, ␴3, and ␴1 are activated for mRNA
synthesis (3, 21, 37, 60). Derepression of the core-associated
transcriptases correlates with protease digestion of ␮1, and a
difference among strains in the capacity of particles to undergo
transcriptase activation is genetically determined by M2 (11,
21, 37). Conversely, addition of ␮1 and ␴3 to cores represses
mRNA synthesis (24). These observations strongly suggest that
the outer capsid in general, and protein ␮1 in particular, play
a role in regulating the activity of the core-associated transcriptases. Other experiments showed that incubation of ISVPs with
K⫹ or Cs⫹ ions at 37°C is sufficient for derepression of transcriptase activity and that proteolytic removal of ␮1 from viral
particles is not required (9, 12, 23, 57). In the present study, we
found that incubation of ISVPs with Cs⫹ ions accelerates, in
concert, both conformational change in ␮1 (Fig. 3) and loss of
␴1 from particles (Fig. 7 and 8). Putting previous and present
findings together, we speculated that one or more structural
changes in ISVPs identified in this study may also play a role in
transcriptase activation. To test that hypothesis, T1L virions or
ISVPs were incubated with Na⫹ or Cs⫹ ions at 32°C and then
removed to ice. Samples were incubated at 37°C in a transcription reaction mixture containing all four ribonucleoside
triphosphates, [␣-32P]GTP, Mg2⫹ ions, and an ATP-regenerating system. Long transcription products were separated from
Downloaded from http://jvi.asm.org/ on February 27, 2014 by PENN STATE UNIV
FIG. 8. Effect of ␮1 conformational change on particle association
of protein ␴1. (A) Nonpurified T1L ISVPs (4 ⫻ 1012 particles/ml) were
incubated in reaction buffer containing NaCl or CsCl (300 mM) for 20
min at 32°C and then removed to ice. Each sample was overlaid onto
a preformed CsCl gradient (␳ ⫽ 1.30 to 1.45 g/cm3, 3.2 ml) and
subjected to centrifugation in a Beckman SW60 rotor (30,000 rpm for
2 h at 5°C). Gradients were photographed with a Nikon Coolpix digital
camera. Images were cropped and resized in Photoshop, version 5.5
(Adobe Systems, San Jose, Calif.). Arrowhead, position of the band in
each gradient. (B) The banded material in panel A was concentrated
by TCA precipitation and subjected to SDS-PAGE, followed by Coomassie staining to visualize the viral proteins (top) or immunoblot
analysis to visualize ␴1. ⴱ, position of an unidentified proteolytic fragment that comigrated with ␴1. The conformational state of ␮1 in each
sample was assessed by trypsin treatment and is indicated below the
gels and immunoblots. (C) Nonpurified T1L ISVPs (4 ⫻ 1012 particles/
ml) were incubated in reaction buffer containing NaCl or CsCl (300
mM) for different times at 32°C and then removed to ice. Each sample
was overlaid onto a sucrose cushion (20% [wt/vol], 500 ␮l) and subjected to centrifugation in a Beckman TLA 100.2 rotor (90,000 rpm for
1 h at 5°C). A 200-␮l fraction was removed from the top of each
gradient, and the proteins in this fraction were concentrated by TCA
precipitation and subjected to SDS-PAGE. Protein ␴1 was visualized
by immunoblot analysis. The conformational state of ␮1 in each sample was assessed by trypsin treatment and is indicated below each
immunoblot.
J. VIROL.
VOL. 76, 2002
ENTRY-RELATED STRUCTURAL TRANSITION IN REOVIRUS ISVPs
oligonucleotides and free nucleotides by precipitation with
TCA, and 32P incorporation into the precipitated RNA was
quantitated by scintillation counting. Viral particles containing
a protease-resistant conformer of ␮1 (i.e., virions incubated
with Na⫹ or Cs⫹ ions and ISVPs incubated with Na⫹ ions)
remained inactive at mRNA synthesis (Fig. 9). In contrast,
particles containing a protease-sensitive conformer of ␮1 (i.e.,
ISVPs incubated with Cs⫹ ions) were active for mRNA synthesis (Fig. 9). These results suggest that ␮1 conformational
change or loss of ␴1 from particles or both are necessary for
derepression of the core-associated viral transcriptases.
DISCUSSION
In this report we present evidence that reovirus ISVPs can
undergo a structural transition to a distinct particle type, the
ISVP*, when incubated at physiological temperature. We show
that ISVP*s contain an altered conformer of the penetration
protein ␮1 (i.e., ␮1*), in which previously buried hydrophobic
regions are now exposed (Fig. 2 to 8 and 10). We also show
that ISVP*s have lost receptor-binding protein ␴1 (Fig. 7, 8,
and 10). Induction of the ISVP-to-ISVP* transition is required
for permeabilization of RBC membranes (Fig. 4) and activation of the core-associated viral transcriptases (Fig. 9). Furthermore, the ␮1* conformer is itself required for hemolytic
activity (Fig. 5). Our findings support an entry model in which
the ISVP per se does not mediate membrane penetration but,
instead, is first converted to the ISVP* form, which then plays
the central role in membrane penetration. We propose that the
␮1 conformational change occurring as part of the ISVP-toISVP* transition is analogous to the entry-related conforma-
tional transitions in enveloped-virus fusion machines, such as
the influenza virus HA and TBEV E proteins, and nonenveloped virus penetration machines, such as the poliovirus VP1
and VP4 proteins (Fig. 10).
Primed and triggered conformational changes in metastable
viral structures: parallels between enveloped viruses and the
nonenveloped reoviruses. Influenza virus HA homotrimers are
synthesized and assembled as HA0 precursors, which are inactive at fusion. HA0 undergoes an endolytic cleavage by a furinlike protease to yield HA1 and HA2. This cleavage primes HA
for fusion by liberating sequences in HA2 for fusion-related
conformational changes (39, 63). Monomers of E protein in
fusion-inactive precursor TBEV particles exist in heterodimeric
complex with protein prM. Proteolytic degradation of a region
of prM by a furin-like protease allows homodimerization of the
E protein, thereby priming E for its conformational change
(32, 33). Priming by proteolytic cleavage at a single site within
the fusion protein itself or more extensive processing of an
interacting protein provides a mechanism to prevent inappropriate induction of conformational changes in the fusion machine that may inactivate viral particles or kill host cells.
The proteolytic degradation of ␴3 and/or more-limited
cleavage(s) of ␮1 by lysosomal proteases (1, 2, 40) may represent analogous priming mechanisms for reovirus particles (Fig.
10). Virions, which contain ␴3 in complex with uncleaved ␮1
homotrimers, do not undergo the ␮1-to-␮1* change (Fig. 9)
and cannot mediate hemolysis (Fig. 1). In contrast, ISVPs,
which lack ␴3 and which contain ␮1 mostly as cleaved fragments, are competent to do both (Fig. 2). The endolytic cleavage of ␮1 at the ␦-␾ junction during conversion of virions to
ISVPs appears to play no part in priming ␮1 (15, 16), however,
suggesting a primary role for ␴3 degradation in this process
(36). How might ␴3 degradation prime ␮1 for its conformational change? Examination of the recently determined crystal
structure of the (␮1)3/(␴3)3 heterohexamer (42) reveals that
each ␴3 monomer interacts extensively with two adjacent subunits in the ␮1 trimer. ␴3 may thus function as a molecular
clamp, inhibiting the conformational freedom of individual ␮1
subunits within the trimer.
Primed HA1-HA2 trimers and E dimers change conformation in response to a low-pH trigger. HA can also be induced
to change conformation by nonphysiological stimuli such as
elevated temperature and pressure (14, 28). For both HA and
E, the conformationally rearranged form of the fusion protein
is more thermostable than its precursor (18, 65). Such observations led to the hypothesis that the primed fusion proteins
are kinetically trapped in a metastable state (14, 65). We show
here that reovirus ISVPs, but not virions, are accelerated to
undergo the ␮1-to-␮1* change by Cs⫹ ions (Fig. 3). The rate of
this change is also increased by other monovalent cations, such
as K⫹, and by elevated particle concentration (57; M. L. Nibert
and K. Chandran, unpublished data). Moreover, we recently
showed that thermal inactivation of ISVPs is associated with a
conformational change in ␮1 that renders it protease sensitive
and that ISVPs are less thermostable than virions (36, 47).
Although the mechanistic bases for the accelerating effects of
monovalent cations, particle concentration, and temperature
on ␮1 conformational change are poorly understood, the
above findings lead us to speculate that removal of ␴3 renders
the ␮1 trimers in ISVPs metastable, that is, energetically
Downloaded from http://jvi.asm.org/ on February 27, 2014 by PENN STATE UNIV
FIG. 9. Effect on ␮1 conformational change on activity of the particle-associated viral transcriptases. Purified T1L virions and nonpurified T1L ISVPs (4 ⫻ 1012 particles/ml) were incubated in reaction
buffer containing NaCl or CsCl (300 mM) at 32°C for 20 min and then
removed to ice. Samples were then incubated with a transcription
reaction mixture including ribonucleoside triphosphates and
[␣-32P]GTP for 1 h at 37°C as described previously (44). The amount
of 32P incorporated into reovirus mRNA was measured by TCA precipitation followed by liquid scintillation counting (44). Averages ⫾
standard deviations of three trials are shown. The conformational state
of ␮1 in each sample was assessed by trypsin treatment and is indicated
below the graph.
9929
9930
CHANDRAN ET AL.
J. VIROL.
poised to undergo a conformational change in response to a
triggering stimulus (Fig. 10). The identity of the physiological
trigger for the ␮1-to-␮1* change is not yet known (but see
below).
Changes in ␮1 structure. ␮1 folds into four domains in
ISVPs. Domains I to III are mainly ␣-helical in nature and
form the intertwined body of the ␮1 trimer. Domain IV is a
jelly roll ␤-barrel that constitutes the top domain of each subunit within the trimer (42). The ␦ region makes up portions of
domains I to III and domain IV in its entirety. Our observation
that ␦ is extensively protease sensitive in ␮1* (Fig. 2 and 3)
indicates that sequences within this region have undergone
refolding during the ␮1-to-␮1* transition. The results of partial-proteolysis experiments suggest that most or all of these
refolding events occur in the lower domains (I to III) and not
in the top domain (IV) (35; M. L. Nibert and K. Chandran,
unpublished data). Domains I to III contain several hydropho-
bic sequences and long amphipathic ␣-helices that may become reorganized during ␮1 conformational change and that
may account for the hydrophobicity of ␮1* as well as the
putative ␮1*-membrane interactions that lead to hemolysis.
Most of these sequences are buried within the ␮1 trimer, and
partial or total disruption of intersubunit interactions may be
necessary to permit their exposure.
Although the present study focused on changes in the central ␦ region of ␮1, the N- and/or C-terminal regions of ␮1
(␮1N and ␾, respectively) may also rearrange as part of the ␮1
conformational change. The highly apolar ␮1N region (Nmyristoyl group and 41 amino acids), a prime candidate for
mediating ␮1*-membrane interactions, is buried within the
base of the ␮1 trimer in ISVPs (42). Its exposure during the
␮1-to-␮1* change and the subsequent insertion into a membrane seem to require not only the reorganization of sequences
from the ␦ region that sandwich ␮1N in the trimer but also
Downloaded from http://jvi.asm.org/ on February 27, 2014 by PENN STATE UNIV
FIG. 10. Parallels between the membrane penetration machines of enveloped and nonenveloped viruses. It is proposed that the membrane
penetration protein of the nonenveloped reoviruses (E) resembles the membrane fusion proteins of enveloped viruses such as influenza virus
(B) and TBEV (C) as well as the membrane penetration proteins of other nonenveloped viruses such as poliovirus (D). All of these proteins appear
to undergo primed and triggered rearrangements that yield a hydrophobic protein conformer capable of interacting with membranes. Proteinmembrane interactions have different consequences for the enveloped and nonenveloped viruses: membrane fusion for the former and either
membrane pore formation or membrane perforation for the latter. Some essential structural and functional properties of viral particles containing
each type of protein conformer are summarized in panel A. (E) Schematic of radial sections of reovirus particle forms containing the different
conformers of membrane penetration protein ␮1. Also indicated are other changes that accompany the structural transitions in reovirus particles:
␴3 degradation (virion to ISVP) and ␴1 elution and ␭2 conformational change (ISVP to ISVP*). Proteins are colored as in Fig. 1A.
VOL. 76, 2002
ENTRY-RELATED STRUCTURAL TRANSITION IN REOVIRUS ISVPs
mediate attachment by binding to cell surface receptors (4, 6).
Interestingly, in addition to their structural and functional similarities, the Ad fiber and ␴1 may have similar fates during the
entry-related disassembly of their respective particles. The Ad
fiber dissociates from particles after viral attachment but before penetration (31, 48), and ␴1 loss occurs in concert with
activation of particles for membrane permeabilization (Fig. 8).
We propose that ␴1, like the Ad fiber, dissociates from particles after attachment to cells but before or during the membrane penetration step.
Triggering the ISVP-to-ISVP* transition. Loss of the fiber
protein from Ad virions is thought to require a conformational
change in the penton base, triggered by its binding to the
integrin coreceptor (31, 48). Our observation that T1L ISVPs
undergo ␴1 loss and the ␮1-to-␮1* change only slowly at physiological temperature (Fig. 2 and 6 to 8) suggests that a triggering stimulus is also required to induce the ISVP-to-ISVP*
transition within cells. One possibility is that the binding of ␴1
to the receptor induces the opening of the ␭2 turret and ␴1
loss. The ␭2 conformational change may then propagate to ␮1
and trigger it to change conformation. In support of this idea,
it has been suggested that reovirus attachment involves conformational changes in ␴1 that propagate to other outer-capsid
proteins (17, 25). A ␴1-receptor interaction is unlikely to be
the sole trigger for the ISVP-to-ISVP* transition, however,
since viral particles can undergo this transition in the absence
of the receptor (Fig. 3) or ␴1 (16; K. Chandran and M. L.
Nibert, unpublished data). We favor, instead, a model in which
the ␮1-to-␮1* change initiates the ISVP-to-ISVP* transition.
The ␮1 conformational change then induces constitutive or
transient opening of the ␭2 turret, which in turn results in ␴1
release. Consistent with this alternative model is our previous
observation that addition of ␮1, but not ␴1, to cores induces
the reverse conformational change, namely, closure of the ␭2
turret (16). The ␮1-to-␮1* transition may be triggered during
entry by K⫹ ions (57) (data not shown), which, like Cs⫹ ions,
accelerate the transition in vitro (Fig. 3). It is also conceivable
that direct interaction of ␮1 with a host cell factor such as a
lipid or protein is the inducing signal.
Contributions of ␮1, ␴1, and ␭2 to membrane penetration
and derepression of the viral transcriptases. Our present findings indicate that ␮1 plays the major role in membrane penetration but do not exclude participation of the other outercapsid proteins. The following observations, however, suggest
that involvement of ␴1 and/or ␭2 is unlikely to take the form of
protein interactions with the membrane bilayer. (i) The hydrophobicity of ␴1 does not increase in conjunction with its dissociation from particles (Fig. 7); (ii) ␴1 affects neither the
extent nor the rate of hemolysis (16; K. Chandran and M. L.
Nibert, unpublished data); and (iii) cores, which lack ␮1 and
contain a more open conformer of the ␭2 turret, do not lyse
RBCs (Fig. 1) (16). An alternative possibility is that ␴1 and ␭2
participate not by interacting with the target membrane but by
facilitating ␮1*-membrane interactions. For example, the
opening of the ␭2 turret followed by release of ␴1 during the
ISVP-to-ISVP* transition may be necessary to allow ␮1* to
approach the membrane. In addition, loss of ␴1 may be necessary to disengage infecting particles from receptors and permit their release into the cytoplasm after penetration. We
Downloaded from http://jvi.asm.org/ on February 27, 2014 by PENN STATE UNIV
autolytic cleavage at the ␮1N-␦ junction (42, 53; A. L. Odegard
and M. L. Nibert, unpublished data). The nature of changes in
␮1 tertiary and quaternary structure during the ␮1-to-␮1*
transition, as well as identities of the ␮1* sequences that insert
into membrane, are currently unknown but are topics of ongoing investigation in our laboratory.
Common features of nonenveloped-virus penetration proteins. It has been proposed that externalization of a buried
N-myristoylated peptide plays a role in membrane penetration
by the nonenveloped picornaviruses and polyomaviruses (19,
34). With poliovirus, associated structural changes include externalization of the hydrophobic N terminus of capsid protein
VP1 (26) and extensive “tectonic” movements throughout the
capsid (5). These conformational changes in poliovirus are
triggered by receptor binding to the capsid surface (30). In
addition, cleavages of the poliovirus capsid polyprotein are
required for these events (34), and the mature capsid is metastable (68). The rotavirus penetration protein (VP4) is not
N-myristoylated but contains an internal hydrophobic sequence, similar to a fusion peptide, near which cleavage must
occur before the protein is active (29, 49). The Ad penton base
protein is neither N-myristoylated nor cleaved; moreover, conformational changes in this protein appear to occur but are not
well characterized (58). Thus, although few if any features have
been shown to be consistently shared by the penetration proteins of all nonenveloped viruses, some common features include N-myristoylation, priming by proteolytic cleavage(s),
metastability of the primed proteins, and triggered conformational changes with exposure of the myristoyl group and/or
other hydrophobic regions in association with membrane interaction (Fig. 10). The lack of uniformly consistent features
may reflect the fact that several variations to the membrane
penetration mechanism are employed by the different nonenveloped viruses.
Loss of ␴1 from particles and putative conformational
change in ␭2. Our finding that ␴1 loss and ␮1 conformational
change are temporally correlated (Fig. 8) suggests that the two
events are mechanistically linked. Since these proteins are
thought not to contact each other in particles, ␴1 loss may be
coupled to ␮1 conformational change via a third protein that
interacts with both of them. Several observations implicate ␭2
as this protein. First, ␴1 binds at the center of the pentameric
“shutter” atop the ␭2 turret and ␮1 subunits make extensive
contacts with the walls of the ␭2 turret around each fivefold
axis (22). Second, when ISVPs are digested to cores in vitro, ␭2
becomes capable of opening the pentameric shutter (22).
Third, ␴1 cannot stably bind to this more open conformer of
the turret (K. Chandran and M. L. Nibert, unpublished data).
We thus conclude that the ␭2 turret opens either constitutively
or transiently during the ISVP-to-ISVP* transition and that ␴1
is lost from viral particles as a consequence of this change in
the turret.
The recently determined structure of a portion of ␴1 revealed a striking resemblance to the Ad fiber protein (17, 69).
Both proteins form trimers comprising a fibrous tail and a
globular head. Both tails are composed in part of a ␤-spiral
motif, and both heads contain eight-stranded ␤-barrel domains. Ad fiber and ␴1 are both bound to a pentameric structure at the fivefold axes of a viral particle, the penton base for
the Ad fiber (64) and the ␭2 turret for ␴1 (22). Both proteins
9931
9932
CHANDRAN ET AL.
ACKNOWLEDGMENTS
We thank L. A. Breun and E. Freimont for excellent technical
support. Many thanks also go to the other members of our laboratory,
S. Liemann, and S. C. Harrison for helpful discussions and to M.
Agosto, S. C. Harrison, K. S. Myers, A. L. Odegard, and J. S. L. Parker
for critical reviews of the manuscript.
This work was supported by NIH grants R29 AI39533 and R01
AI46440 (to M.L.N.). K.C. was additionally supported by a predoctoral
fellowship from the Howard Hughes Medical Institute and a Fields
postdoctoral fellowship made available to the Department of Microbiology and Molecular Genetics through the generosity of Ruth Peedin
Fields.
REFERENCES
1. Baer, G. S., and T. S. Dermody. 1997. Mutations in reovirus outer-capsid
protein ␴3 selected during persistent infections of L cells confer resistance to
protease inhibitor E64. J. Virol. 71:4921–4928.
2. Baer, G. S., D. H. Ebert, C. J. Chung, A. H. Erickson, and T. S. Dermody.
1999. Mutant cells selected during persistent reovirus infection do not express mature cathepsin L and do not support reovirus disassembly. J. Virol.
73:9532–9543.
3. Banerjee, A. K., and A. J. Shatkin. 1970. Transcription in vitro by reovirusassociated ribonucleic acid-dependent polymerase. J. Virol. 6:1–11.
4. Barton, E. S., J. C. Forrest, J. L. Connolly, J. D. Chappell, Y. Liu, F. J.
Schnell, A. Nusrat, C. A. Parkos, and T. S. Dermody. 2001. Junction adhesion molecule is a receptor for reovirus. Cell 104:441–451.
5. Belnap, D. M., D. J. Filman, B. L. Trus, N. Cheng, F. P. Booy, J. F. Conway,
S. Curry, C. N. Hiremath, S. K. Tsang, A. C. Steven, and J. M. Hogle. 2000.
Molecular tectonic model of virus structural transitions: the putative cell
entry states of poliovirus. J. Virol. 74:1342–1354.
6. Bewley, M. C., K. Springer, Y. B. Zhang, P. Freimuth, and J. M. Flanagan.
1999. Structural analysis of the mechanism of adenovirus binding to its
human cellular receptor, CAR. Science 286:1579–1583.
7. Blumenthal, R., P. Seth, M. C. Willingham, and I. Pastan. 1986. pH dependent lysis of liposomes by adenovirus. Biochemistry 25:2231–2237.
8. Bordier, C. 1981. Phase separation of integral membrane proteins in Triton
X-114 solution. J. Biol. Chem. 256:1604–1607.
9. Borsa, J., D. G. Long, T. P. Copps, M. D. Sargent, and J. D. Chapman. 1974.
Reovirus transcriptase activation in vitro: further studies on the facilitation
phenomenon. Intervirology 3:15–35.
10. Borsa, J., B. D. Morash, M. D. Sargent, T. P. Copps, P. A. Lievaart, and J. G.
Szekely. 1979. Two modes of entry of reovirus particles into L cells. J. Gen.
Virol. 45:161–170.
11. Borsa, J., M. D. Sargent, T. P. Copps, D. G. Long, and J. D. Chapman. 1973.
Specific monovalent cation effects on modification of reovirus infectivity by
chymotrypsin digestion in vitro. J. Virol. 11:1017–1019.
12. Borsa, J., M. D. Sargent, D. D. Ewing, and M. Einspenner. 1982. Perturbation of the switch-on of transcriptase activity in intermediate subviral particles from reovirus. J. Cell. Physiol. 112:10–18.
13. Borsa, J., M. D. Sargent, P. A. Lievaart, and T. P. Copps. 1981. Reovirus:
evidence for a second step in the intracellular uncoating and transcriptase
activation process. Virology 111:191–200.
14. Carr, C. M., C. Chaudhry, and P. S. Kim. 1997. Influenza hemagglutinin is
spring-loaded by a metastable native conformation. Proc. Natl. Acad. Sci.
USA 94:14306–14313.
15. Chandran, K., and M. L. Nibert. 1998. Protease cleavage of reovirus capsid
protein ␮1/␮1C is blocked by alkyl sulfate detergents, yielding a new type of
infectious subvirion particle. J. Virol. 72:467–475.
16. Chandran, K., S. B. Walker, Y. Chen, C. M. Contreras, L. A. Schiff, T. S.
Baker, and M. L. Nibert. 1999. In vitro recoating of reovirus cores with
baculovirus-expressed outer-capsid proteins ␮1 and ␴3. J. Virol. 73:3941–
3950.
17. Chappell, J. D., A. E. Prota, T. S. Dermody, and T. Stehle. 2002. Crystal
structure of reovirus attachment protein ␴1 reveals evolutionary relationship
to adenovirus fiber. EMBO J. 21:1–11.
18. Chen, J., S. A. Wharton, W. Weissenhorn, L. J. Calder, F. M. Hughson, J. J.
Skehel, and D. C. Wiley. 1995. A soluble domain of the membrane-anchoring
chain of influenza virus hemagglutinin (HA2) folds in E. coli into the lowpH-induced conformation. Proc. Natl. Acad. Sci. USA 92:12205–12209.
19. Chen, X. S., T. Stehle, and S. C. Harrison. 1998. Interaction of polyomavirus
internal protein VP2 with the major capsid protein VP1 and implications for
participation of VP2 in viral entry. EMBO J. 17:3233–3240.
20. Coombs, K. M. 1998. Stoichiometry of reovirus structural proteins in virus,
ISVP, and core particles. Virology 243:218–228.
21. Drayna, D., and B. N. Fields. 1982. Activation and characterization of the
reovirus transcriptase: genetic analysis. J. Virol. 41:110–118.
22. Dryden, K. A., G. Wang, M. Yeager, M. L. Nibert, K. M. Coombs, D. B.
Furlong, B. N. Fields, and T. S. Baker. 1993. Early steps in reovirus infection
are associated with dramatic changes in supramolecular structure and protein conformation: analysis of virions and subviral particles by cryoelectron
microscopy and image reconstruction. J. Cell Biol. 122:1023–1041.
23. Ewing, D. D., M. D. Sargent, and J. Borsa. 1985. Switch-on of transcriptase
function in reovirus: analysis of polypeptide changes using 2-D gels. Virology
144:448–456.
24. Farsetta, D. L., K. Chandran, and M. L. Nibert. 2000. Transcriptional activities of reovirus RNA polymerase in recoated cores. Initiation and elongation are regulated by separate mechanisms. J. Biol. Chem. 275:39693–
39701.
25. Fernandes, J., D. Tang, G. Leone, and P. W. Lee. 1994. Binding of reovirus
to receptor leads to conformational changes in viral capsid proteins that are
reversible upon virus detachment. J. Biol. Chem. 269:17043–17047.
26. Fricks, C. E., and J. M. Hogle. 1990. Cell-induced conformational change in
poliovirus: externalization of the amino terminus of VP1 is responsible for
liposome binding. J. Virol. 64:1934–1945.
27. Furlong, D. B., M. L. Nibert, and B. N. Fields. 1988. ␴1 protein of mammalian reoviruses extends from the surfaces of viral particles. J. Virol. 62:246–
256.
28. Gaspar, L. P., A. C. Silva, A. M. Gomes, M. S. Freitas, A. P. Ano Bom, W. D.
Schwarcz, J. Mestecky, M. J. Novak, D. Foguel, and J. L. Silva. 2001.
Hydrostatic pressure induces the fusion-active state of enveloped viruses.
J. Biol. Chem. 277:8433–8439.
29. Gilbert, J. M., and H. B. Greenberg. 1997. Virus-like particle-induced fusion
from without in tissue culture cells: role of outer-layer proteins VP4 and
VP7. J. Virol 71:4555–4563.
30. Gomez Yafal, A., G. Kaplan, V. R. Racaniello, and J. M. Hogle. 1993.
Characterization of poliovirus conformational alteration mediated by soluble
cell receptors. Virology 197:501–505.
31. Greber, U. F. 1998. Virus assembly and disassembly: the adenovirus cysteine
protease as a trigger factor. Rev. Med. Virol. 8:213–222.
32. Guirakhoo, F., F. X. Heinz, C. W. Mandl, H. Holzmann, and C. Kunz. 1991.
Fusion activity of flaviviruses: comparison of mature and immature (prMcontaining) tick-borne encephalitis virions. J. Gen. Virol. 72:1323–1329.
33. Heinz, F. X., and S. L. Allison. 2001. The machinery for flavivirus fusion with
host cell membranes. Curr. Opin. Microbiol. 4:450–455.
34. Hindiyeh, M., Q. H. Li, R. Basavappa, J. M. Hogle, and M. Chow. 1999.
Poliovirus mutants at histidine 195 of VP2 do not cleave VP0 into VP2 and
VP4. J. Virol. 73:9072–9079.
35. Hooper, J. W., and B. N. Fields. 1996. Role of the ␮1 protein in reovirus
Downloaded from http://jvi.asm.org/ on February 27, 2014 by PENN STATE UNIV
speculate that release of Ad fibers from particles may fulfill
similar functions in membrane penetration by that virus.
Derepression of the reovirus core-associated transcriptases
coincides with the ISVP-to-ISVP* transition (Fig. 8). A role
for the ␮1-to-␮1* change in this process is certain, given previous evidence implicating M2 and ␮1 as determinants of both
transcriptase activation and repression (21, 23, 24). How does
␮1* allow derepression of the viral transcription machines
located within the inner capsid? In ISVPs, large loops that
extend from each ␮1 trimer contact the raised nodules composed of inner-capsid protein ␴2 (22, 42, 56). ␴2 may therefore
relay a signal from ␮1* to the transcription machines. Alternatively or in addition, ␮1* may signal the transcriptases by
inducing the ␭2 turret to open. The ␭2 conformational change
may effect activation by propagating to the capsid interior via
inner-capsid protein ␭1 or ␴2 or both. Furthermore, because
viral mRNAs are thought to exit the particle via the central
channel of the ␭2 turret, both the opening of the turret and the
loss of ␴1 may be necessary for unimpeded mRNA egress and
continued transcription (24; M. Yeager, S. Weiner, and K. M.
Coombs, Abstr. 40th Meet. Biophys. Soc., abstr. 116, 1996).
Our hypothesis that ISVP*s are, or closely resemble, primary
transcriptase particles is consistent with previous observations
that the latter resemble ISVPs, and not cores, in protein composition (Fig. 8) (38, 62). Unlike activation of the virus-membrane interaction, however, activation of the core-associated
viral transcriptases requires only the ␮1-to-␮1* change and not
the continued presence of ␮1* (37) (data not shown). Thus,
␮1* may be removed from particles during or after membrane
penetration.
J. VIROL.
VOL. 76, 2002
36.
37.
38.
39.
40.
42.
43.
44.
45.
46.
47.
48.
49.
50.
51.
52.
53.
stability and capacity to cause chromium release from host cells. J. Virol.
70:459–467.
Jané-Valbuena, J., M. L. Nibert, S. M. Spencer, S. B. Walker, T. S. Baker, Y.
Chen, V. E. Centonze, and L. A. Schiff. 1999. Reovirus virion-like particles
obtained by recoating infectious subvirion particles with baculovirus-expressed ␴3 protein: an approach for analyzing ␴3 functions during virus
entry. J. Virol. 73:2963–2973.
Joklik, W. K. 1972. Studies on the effect of chymotrypsin on reovirions.
Virology 49:700–715.
Joklik, W. K., and M. R. Roner. 1995. What reassorts when reovirus genome
segments reassort? J. Biol. Chem. 270:4181–4184.
Klenk, H. D., R. Rott, M. Orlich, and J. Blodorn. 1975. Activation of
influenza A viruses by trypsin treatment. Virology 68:426–439.
Kothandaraman, S., M. C. Hebert, R. T. Raines, and M. L. Nibert. 1998. No
role for pepstatin-A-sensitive acidic proteinases in reovirus infections of L or
MDCK cells. Virology 251:264–272.
Lee, P. W. K., E. C. Hayes, and W. K. Joklik. 1981. Protein ␴1 is the reovirus
cell attachment protein. Virology 108:156–163.
Liemann, S., K. Chandran, T. S. Baker, M. L. Nibert, and S. C. Harrison.
2002. Structure of the reovirus membrane-penetration protein, ␮1, in a
complex with is protector protein, ␴3. Cell 108:283–295.
Lucia-Jandris, P., J. W. Hooper, and B. N. Fields. 1993. Reovirus M2 gene
is associated with chromium release from mouse L cells. J. Virol. 67:5339–
5345.
Luongo, C. L., K. A. Dryden, D. L. Farsetta, R. L. Margraf, T. F. Severson,
N. H. Olson, B. N. Fields, T. S. Baker, and M. L. Nibert. 1997. Localization
of a C-terminal region of ␭2 protein in reovirus cores. J. Virol. 71:8035–8040.
Luongo, C. L., K. M. Reinisch, S. C. Harrison, and M. L. Nibert. 2000.
Identification of the guanylyltransferase region and active site in reovirus
mRNA capping protein ␭2. J. Biol. Chem. 275:2804–2810.
McClure, W. O., and G. M. Edelman. 1966. Fluorescent probes for conformational states of proteins. I. Mechanism of fluorescence of 2-p-toluidinylnaphthalene-6-sulfonate, a hydrophobic probe. Biochemistry 5:1908–1919.
Middleton, J. K., T. F. Severson, K. Chandran, A. L. Gillian, J. Yin, and
M. L. Nibert. 2002. Thermostability of reovirus disassembly intermediates
(ISVPs) correlates with genetic, biochemical, and thermodynamic properties
of major surface protein ␮1. J. Virol. 76:1051–1061.
Nakano, M. Y. B., K. M. Suomalainen, R. P. Stidwill, and U. F. Greber. 2000.
The first step of adenovirus type 2 disassembly occurs at the cell surface,
independently of endocytosis and escape to the cytosol. J. Virol. 74:7085–
7095.
Nandi, P., A. Charpilienne, and J. Cohen. 1992. Interaction of rotavirus
particles with liposomes. J. Virol. 66:3363–3367.
Nibert, M. L. 1998. Structure of mammalian orthoreovirus particles. Curr.
Top. Microbiol. Immunol. 238(Reovirus i):1–30.
Nibert, M. L., and B. N. Fields. 1992. A carboxy-terminal fragment of protein
␮1/␮1C is present in infectious subvirion particles of mammalian reoviruses
and is proposed to have a role in penetration. J. Virol. 66:6408–6418.
Nibert, M. L., and L. A. Schiff. 2001. Reoviruses and their replication, p.
1679–1728. In D. M. Knipe and P. M. Howley (ed.), Fields virology, 4th ed.
Lippincott Williams & Wilkins, Philadelphia, Pa.
Nibert, M. L., L. A. Schiff, and B. N. Fields. 1991. Mammalian reoviruses
contain a myristoylated structural protein. J. Virol. 65:1960–1967.
9933
54. Noble, S., and M. L. Nibert. 1997. Characterization of an ATPase activity in
reovirus cores and its genetic association with core-shell protein ␭1. J. Virol.
71:2182–2191.
55. Parker, M. W., A. D. Tucker, D. Tsernoglou, and F. Pattus. 1990. Insights
into membrane insertion based on studies of colicins. Trends Biochem. Sci.
15:126–129.
56. Reinisch, K. M., M. L. Nibert, and S. C. Harrison. 2000. Structure of the
reovirus core at 3.6 Å resolution. Nature 404:960–967.
57. Sargent, M. D., D. G. Long, and J. Borsa. 1977. Functional analysis of the
interactions between reovirus particles and various proteases in vitro. Virology 78:354–358.
58. Seth, P., M. C. Willingham, and I. Pastan. 1985. Binding of adenovirus and
its external proteins to Triton X-114. J. Biol. Chem. 260:14431–14434.
59. Shai, Y. 1999. Mechanism of the binding, insertion and destabilization of
phospholipid bilayer membranes by alpha-helical antimicrobial and cell nonselective membrane-lytic peptides. Biochim. Biophys. Acta 1462:55–70.
60. Shatkin, A. J., and J. D. Sipe. 1968. RNA polymerase activity in purified
reoviruses. Proc. Natl. Acad. Sci. USA 61:1462–1469.
61. Shepard, D. A., J. G. Ehnstrom, and L. A. Schiff. 1995. Association of
reovirus outer capsid proteins ␴3 and ␮1 causes a conformational change
that renders ␴3 protease sensitive. J. Virol. 69:8180–8184.
62. Silverstein, S. C., C. Astell, D. H. Levin, M. Schonberg, and G. Acs. 1972.
The mechanisms of reovirus uncoating and gene activation in vivo. Virology
47:797–806.
63. Skehel, J. J., and D. C. Wiley. 2000. Receptor binding and membrane fusion
in virus entry: the influenza hemagglutinin. Annu. Rev. Biochem. 69:531–
569.
64. Stewart, P. L., R. M. Burnett, M. Cyrklaff, and S. D. Fuller. 1991. Image
reconstruction reveals the complex molecular organization of adenovirus.
Cell 67:145–154.
65. Stiasny, K., S. L. Allison, C. W. Mandl, and F. X. Heinz. 2001. Role of
metastability and acidic pH in membrane fusion by tick-borne encephalitis
virus. J. Virol. 75:7392–7398.
66. Sturzenbecker, L. J., M. Nibert, D. Furlong, and B. N. Fields. 1987. Intracellular digestion of reovirus particles requires a low pH and is an essential
step in the viral infectious cycle. J. Virol. 61:2351–2361.
67. Tosteson, M. T., M. L. Nibert, and B. N. Fields. 1993. Ion channels induced
in lipid bilayers by subvirion particles of the nonenveloped mammalian
reoviruses. Proc. Natl. Acad. Sci. USA 90:10549–10552.
68. Tsang, S. K., P. Danthi, M. Chow, and J. M. Hogle. 2000. Stabilization of
poliovirus by capsid-binding antiviral drugs is due to entropic effects. J. Mol.
Biol. 296:335–340.
69. van Raaij, M. J., A. Mitraki, G. Lavigne, and S. Cusack. 1999. A triple
␤-spiral in the adenovirus fiber shaft reveals a new structural motif for a
fibrous protein. Nature 401:935–938.
70. Virgin, H. W., IV, M. A. Mann, B. N. Fields, and K. L. Tyler. 1991. Monoclonal antibodies to reovirus reveal structure/function relationships between
capsid proteins and genetics of susceptibility to antibody action. J. Virol.
65:6772–6781.
71. Werck-Reichhart, D., I. Benveniste, H. Teutsch, F. Durst, and B. Gabriac.
1991. Glycerol allows low-temperature phase separation of membrane proteins solubilized in Triton X-114: application to the purification of plant
cytochromes P-450 and b5. Anal. Biochem. 197:125–131.
Downloaded from http://jvi.asm.org/ on February 27, 2014 by PENN STATE UNIV
41.
ENTRY-RELATED STRUCTURAL TRANSITION IN REOVIRUS ISVPs